evaluating garch models

Upload: sariciv

Post on 06-Apr-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Evaluating GARCH Models

    1/19

    Journal of Econometrics 110 (2002) 417435

    www.elsevier.com/locate/econbase

    Evaluating GARCH models

    Stefan Lundbergh, Timo Terasvirta

    Department of Economic Statistics, Stockholm School of Economics, P.O. Box 6501,

    SE-113 83 Stockholm, Sweden

    Abstract

    In this paper, a unied framework for testing the adequacy of an estimated GARCH model is

    presented. Parametric Lagrange multiplier (LM) or LM type tests of no ARCH in standardized

    errors, linearity, and parameter constancy are proposed. The asymptotic null distributions of the

    tests are standard, which makes application easy. Versions of the tests that are robust against

    nonnormal errors are provided. The nite sample properties of the test statistics are investigated

    by simulation. The robust tests prove superior to the nonrobust ones when the errors are non-

    normal. They also compare favourably in terms of power with misspecication tests previously

    proposed in the literature.c 2002 Elsevier Science B.V. All rights reserved.

    JEL classication: C22; C52

    Keywords: Conditional heteroskedasticity; Model misspecication test; Nonlinear time series; Parameter

    constancy; Smooth transition GARCH

    1. Introduction

    When modelling the conditional mean, at least when it is a linear function of para-

    meters, the estimated model is regularly subjected to a battery of misspecication tests

    to check its adequacy. The hypothesis of no (conditional) heteroskedasticity, no error

    autocorrelation, linearity, and parameter constancy, to name a few, are tested using

    various methods. As for models of conditional variance, such as the popular GARCH

    model, testing the adequacy of the estimated model has been much less common in

    practice. But then, misspecication tests do exist for GARCH models as well. For ex-

    ample, Bollerslev (1986) already suggested a score or Lagrange multiplier (LM) test

    for testing a GARCH model against a higher order GARCH model. Li and Mak (1994)

    Corresponding author.

    E-mail address: [email protected] (T. Terasvirta).

    0304-4076/02/$ - see front matter c 2002 Elsevier Science B.V. All rights reserved.

    PII: S 0 3 0 4 - 4 0 7 6 ( 0 2 ) 0 0 0 9 6 - 9

  • 8/3/2019 Evaluating GARCH Models

    2/19

    418 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    derived a portmanteau type test for testing the adequacy of a GARCH model. Engle and

    Ng (1993) considered testing the GARCH specication against asymmetry using the

    so-called sign-bias and size-bias tests. Chu (1995) derived a test of parameter constancy

    against a single structural break. This test has a nonstandard asymptotic null distribu-tion, but Chu provided tables for critical values. Recently, Lin and Yang (1999) derived

    another test against a single structural break, based on empirical distribution functions.

    In this article, we present a number of simple misspecication tests for GARCH

    models. The idea is to make misspecication testing easy without sacricing power.

    We consider testing the null of no ARCH in the standardized errors, linearity or sym-

    metry against a smooth transition GARCH, and parameter constancy against smoothly

    changing parameters. A single structural break is nested in the alternative hypothesis

    of our parameter constancy test. A two-regime asymmetric GARCH model such as

    the so-called GJR model (Glosten et al., 1993) is nested in the alternative of smooth

    transition GARCH. Note that the LM test of Bollerslev (1986) ts into our framework.

    Furthermore, we show that under normality the portmanteau test of Li and Mak (1994)

    is asymptotically equivalent to our test of no remaining ARCH. All our tests are LM

    or LM type tests and require only standard asymptotic distribution theory. They may

    be obtained from the same root by merely changing the denitions of the elements

    of the score vector corresponding to the alternative hypothesis.

    Very often in applications, the assumption of a normal error distribution of a GARCH

    process is too restrictive. A useful feature of our tests is that robustifying them against

    nonnormal errors is straightforward. In fact, our Monte Carlo experiments suggest that

    one should always apply the robust versions of our tests. At the sample size 1000, usedin our simulation study and rather typical for GARCH applications, the eciency loss

    compared to the nonrobust tests appears to be minimal when the errors are normal.

    On the other hand, the same simulations show that the nonrobust tests tend to be

    undersized if the error distribution is leptokurtic and that, in this situation, robust

    tests have clearly better power than the nonrobust ones. In terms of power, the tests

    proposed in this paper compare favourably with corresponding tests currently available

    in the literature. This is true for all three cases we consider: no remaining ARCH,

    linearity, and parameter constancy.

    The plan of the paper is as follows. In Section 2, we dene the model. In Section

    3, we discuss testing the null of no ARCH in the standardized errors and compare ourLM-test with the portmanteau test of Li and Mak (1994). Section 4 considers testing

    null hypotheses of linearity and parameter constancy. Section 5 contains results of a

    simulation experiment and Section 6 concludes.

    2. The model

    Consider a time series model with the following structure:

    yt = f(wt;D) + t; (2.1)where f is at least twice continuously dierentiable with respect to D , for all wt=(yt1; u

    t) with yt1 = (1; yt1; : : : ; ytn)

    Rn+1 and exogenous ut= (u1t; : : : ; ukt) Rk,

  • 8/3/2019 Evaluating GARCH Models

    3/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 419

    everywhere in . The error is parameterized as

    t = th1=2t ; (2.2)

    where {t} is a sequence of independent identically distributed random variables withmean zero, unit variance and E3t = 0. Furthermore, the conditional variance ht = W

    stsuch that st = (1;

    2t1; : : : ;

    2tq; ht1; : : : ; htp)

    and W = (0; 1; : : : ; q; 1; : : : ; p) with

    0 0, whereas 1; : : : ; q; 1; : : : ; p satisfy the conditions in Nelson and Cao (1992)

    that ensure the positivity of ht. These conditions allow some of the parameters to be

    negative, unless p = q = 1. Model (2.2) is thus the standard GARCH(p; q) model.

    We assume regularity conditions hold such that the central limit theorem and the

    law of large numbers apply whenever required. For such conditions in the multi-

    variate GARCH(p; q) case, see Comte and Lieberman (2000). In the univariate case,

    their conditions require the density of t

    to be absolutely continuous with respect

    to the Lebesgue measure and positive in a neighbourhood of the origin. Further-

    more, it is required that E8t , which of course implies further restrictions on thedensity of t.

    The assumption E3t = 0 that Comte and Lieberman (2000) do not need guarantees

    block diagonality of the information matrix of the log-likelihood function. However,

    it is not just a technical simplication. We shall consider, among other things, a test

    that has power against asymmetric response to shocks. In deriving such a test it is

    appropriate to assume that the conditional distribution of t given ht, is not skewed.

    3. Testing the adequacy of the GARCH model

    3.1. Testing the null hypothesis of no remaining ARCH

    In order to consider the adequacy of the GARCH model, we formulate a parametric

    alternative to the model. Assume that in (2.2),

    t = ztg1=2t ; (3.1)

    where {zt} is a sequence of independent, identically distributed random variableswith zero mean, unit variance and Ez

    3t = 0. Furthermore, gt = 1 + ?

    vt, where vt =(2t1; : : : ;

    2tm)

    and ? = (1; : : : ; m); j 0; j = 1; : : : ; m. Eq. (3.1) may thus be

    written as

    t = zt(htgt)1=2 (3.2)

    and could be called an ARCH nested in GARCH model as 2tj = 2tj=htj ; j =

    1; : : : ; m. We want to test H0: ? = 0 against ?= 0 and thus follow the standard prac-tice of choosing a two-sided alternative although the elements of ? are constrained

    to be non-negative. Under this hypothesis, gt 1, and the model collapses into aGARCH(p; q) model. (For ways of testing H0: ? = 0 against ? 0 when ht

    0, see

    Lee and King (1993) and Demos and Sentana (1998).)We introduce the following notation. Let t and ht be the error t and the con-

    ditional variance ht, respectively, estimated under H0, and 2

    t = 2t=ht. Furthermore,

  • 8/3/2019 Evaluating GARCH Models

    4/19

    420 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    let xt = h1

    t 9ht=9W (9ht=9W

    denotes 9ht=9W estimated under H0) and vt = (

    2

    t1; : : : ;

    2

    tm). The quasi maximum likelihood approach leads to the following result:

    Theorem 3.1. Consider the model (3.2) where gt = 1 + ?vt and {zt} is a sequence

    of independent identically distributed random variables with zero mean; unit variance

    and Ez3t = 0. Under H0: ? = 0; the statistic

    LM = (1=4T)

    T

    t=1

    (2t=ht 1)vt

    V(W)1

    T

    t=1

    (2t=ht 1)vt

    ; (3.3)

    where

    V(W)1 = (4T=k)

    Tt=1

    vtv

    t T

    t=1

    vtx

    t

    T

    t=1

    xtx

    t

    1 Tt=1

    xtv

    t

    1

    with k=(1=T)T

    t=1(2t=ht1)2 is a consistent estimator of the inverse of the covariance

    matrix of the partial score under the null hypothesis; has an asymptotic 2 distribution

    with m degrees of freedom.

    Proof. See the appendix.

    The test may also be carried out using an articial regression as follows.

    1. Estimate the parameters of the conditional variance model under the null, compute

    2t=ht 1; t= 1; : : : ; T , and SSR0 =

    Tt=1(

    2t=ht 1)2.

    2. Regress (2t=ht 1) on xt; vt and compute the sum of squared residuals, SSR1.

    3. Compute the test statistic LM = T(SSR0 SSR1)=SSR0 or the F-version:

    F =(SSR0 SSR1)=m

    SSR1=(T

    p

    q

    1

    m)

    :

    For the sample sizes relevant in GARCH modelling, there is no essential dierence

    between the properties of LM and F.

    If the standard GARCH parameterization is assumed to be the true data-generating

    process, then it is well-known in practice that the iid error process often cannot be

    assumed normal, although the likelihood is constructed under the assumption of nor-

    mality. It is therefore desirable to make the test robust against nonnormal errors. The

    robust version of the LM test can be constructed following Wooldridge (1991, Proce-

    dure 4.1). The test is carried out as follows:

    1. Obtain the quasi maximum likelihood estimate for W under H0, compute 2t=ht1; vtand xt; t= 1; : : : ; T .

    2. Regress vt on xt, and compute the (m 1) residual vectors rt; t= 1; : : : ; T .

  • 8/3/2019 Evaluating GARCH Models

    5/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 421

    3. Regress 1 on (2t=ht 1)rt and compute the residual sum of squares SSR from that

    regression. The test statistic

    LMR = T SSRhas an asymptotic 2 distribution with m degrees-of-freedom under the null

    hypothesis.

    The purpose of the second step is to purge the eects of the normality assumption

    from vt. The resulting test statistic has the same asymptotic distribution as (3.3) and is,

    under normality, asymptotically ecient; see Wooldridge (1991). Any other consistent

    estimator of W may be employed in place of the quasi-maximum likelihood one.

    3.2. A portmanteau test and a comparison

    Li and Mak (1994) recently introduced a portmanteau statistic for testing the ade-

    quacy of the standard GARCH(p; q) model. The null hypothesis is that the squared and

    standardized error process is not autocorrelated. Let r = (r1; : : : ; r m) be the m1 vector

    of the rst m autocorrelations of {2t=ht} so that H0:r = 0. Assuming normal errors,Li and Mak (1994) showed that under this hypothesis

    Tr is asymptotically normally

    distributed where T is the number of observations and r = (1=2T)

    (2t=ht 1)vt witht= ytf(wt; D); ht = ht(wt; D; W) and vt = (2t1=ht1 1; : : : ; 2tm=htm 1). Further-more, under the null hypothesis the asymptotic covariance matrix of Tr has the form

    Vr(W) = Im Xr(W)G1(W)Xr(W) (3.4)

    as plimTT1

    vt v

    t = 2Im under H0, and Xr(W) = (1=2T)

    xtv

    t . In (3.4),

    G1(W) is a consistent estimator of the relevant block of the information matrix, eval-

    uated at W = W. The portmanteau statistic becomes

    Q(m) = TrVr(W)

    1r (3.5)

    which, as Li and Mak (1994) showed, is asymptotically 2-distributed with m degrees

    of freedom under the null hypothesis. In order to link (3.5) to our theory in the case

    of normal errors, dene

    Q(m) =1

    4T

    (

    2t=ht 1)v

    t

    Vr (W)

    1

    vt (2t=ht 1)

    ; (3.6)

    where Vr (W) = ( 1=2T)(

    vt v

    t

    vt x

    t(

    xtx

    t)1

    xtv

    t ). The only dierence be-

    tween (3.5) and (3.6) is the choice of the consistent estimator of the covariance matrix.

    Vr (W) contains the matrix T1

    vt v

    t whereas its expectation 2Im appears in Vr(W).

    The two covariance matrices are thus asymptotically equal so that the statistics (3.5)

    and (3.6) and thus (3.3) and (3.5) have the same asymptotic null distribution. Thereason for introducing (3.3) as Li and Mak (1994) already derived the same test is

    that obtaining the test using the LM principle makes it easy to see how the test can be

  • 8/3/2019 Evaluating GARCH Models

    6/19

    422 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    made robust against nonnormal errors. Furthermore, the present test is can be viewed

    as a natural extension of Engles (1982) classic LM test of no ARCH to this situation.

    It is well-known that the McLeod and Li (1983) portmanteau test and the Engle (1982)

    test are asymptotically equivalent, see, for example, Luukkonen et al. (1988b). Whenht = 0 in (2.2), the Li and Mak portmanteau test and our LM test collapse into the

    McLeod and Li (1983) and the Engle test, respectively.

    Our test or the Li and Mak test are in principle general misspecication tests but,

    as is clear from above, they also have an LM interpretation, that is, they are LM tests

    against a certain parametric alternative. If the model builder testing his or her GARCH

    model believes that the only relevant alternative is a higher order GARCH model,

    then the LM test of Bollerslev (1986) is an obvious test to use. If a more general

    misspecication of the model is not excluded a priori, then the test of this section can

    be viewed as a useful complement to Bollerslevs test. Other types of misspecication

    will be considered next.

    4. Misspecication of structure

    In this section, we present two misspecication tests for an estimated conditional

    variance model. The rst one is a test against nonlinearity or, in some cases, asym-

    metry. It is a modication of a test in Hagerud (1997). Second, we propose a test of

    parameter constancy against smooth continuous change in parameters. These tests may

    be viewed as conditional variance counterparts of the tests for the nonlinear conditionalmean presented in Eitrheim and Terasvirta (1996). To describe the common features

    in these tests we rst introduce some notation and thereafter consider the two tests

    separately.

    4.1. Notation

    Consider now an augmented version of model (2.2),

    t = zt(ht + gt)1=2; (4.1)

    where gt = g(?1?2; W; st) such that g(0; ?2; W; st) 0, and st is a vector of randomvariables. The null hypothesis to be tested, using this notation, is ?1 =0. As an example,

    the alternative in the LM test of Bollerslev (1986) for testing a GARCH( p; q) model

    against GARCH(p; q + r) or GARCH(p + r; q); r 0, alternatives, belongs to class

    (4.1). As another example, one could think of a Ramsey-type functional form LM

    test, in which case st = st, and gt = 1(Wst)

    2 or gt = (Wst)[exp{1(Wst)} 1], where

    1 is a scalar and 2 = 0. The null hypothesis would, as above, be 1 = 0. This would

    lead to a one-degree-of-freedom test. When applying GARCH models the degrees of

    freedom are not usually a problem, however, so that we may consider more richly parameterized alternatives. We shall concentrate on tests of linearity and parameter

    constancy of the GARCH(p; q) model.

  • 8/3/2019 Evaluating GARCH Models

    7/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 423

    4.2. Testing linearity

    When volatility in return series is modelled with GARCH models, we may sometimes

    expect the response to be a function not only of the size of the shock but also of thedirection. Engle and Ng (1993), see also references therein, considered this possibility.

    We call such a response to a shock asymmetric and parameterize it by generalizing

    the GJRGARCH model of Glosten et al. (1993). This is done in three ways. First,

    we make the transition between the extreme regimes smooth. Second, we incorporate a

    nonlinear version of the quadratic GARCH model of Sentana (1995) in our alternative.

    Finally, while the GJRGARCH model is asymmetric, the present generalization may

    in some special cases remain symmetric, although the model is nonlinear. For smooth

    transition GARCH, see also Hagerud (1997), Gonzalez-Rivera (1998) and Anderson et

    al. (1999). Our maintained model is in fact a special case of the model of Anderson

    et al. (1999). Let

    Hn(xt; ; c) =

    1 + exp

    nl=1

    (xt cl)

    1

    ; 0; c16 6 cn; (4.2)

    where xt is the transition variable at time t, is a slope parameter, and c = (c1; : : : ; cn)

    a location vector. Conditions 0 and c16 6 cn are identifying conditions. When = 0; Hn(xt; ; c) 12 . Typically in practice, n =1 or n =2. The former choice yields astandard logistic function. When the slope parameter , (4.2) with n =1 becomesa step function whose value equals one for xt

    c1 and zero otherwise. When n = 2,(4.2) is symmetric about (c1 + c2)=2, its minimum value, achieved at this point, lies

    between zero and 12

    , and the value of the function approaches unity as xt . When , function (4.2) becomes a double step function that obtains value zero forc16xt c2 and unity otherwise. The logistic function (4.2) is used for parameterizing

    the maintained model. Let Hn = Hn 12 . This transformation simplies notation inderiving the test but does not aect the generality of the argument. The alternative to

    standard GARCH may now be written as (4.1) where

    gt =

    q

    j=1

    0j Hn(tj; ; c) +

    q

    j=1

    1j Hn(tj; ; c)2tj : (4.3)

    Using previous notation, ?1 = and ?2 = (c1; : : : ; cn; 01; : : : ; 0q; 11; : : : ; 1q) in gt,

    and st=st. The sucient but not necessary conditions for ht+gt 0 are 0 0;q

    i=1 0i 0; j 0; j + 1j 0; j = 1; : : : ; q. Assuming n = 1; 0j = 0; j = 1; : : : ; q, and letting

    in (4.3) yields the GJRGARCH model. The test of the standard GARCHmodel against nonlinear GARCH in Hagerud (1997) may be viewed as a special case

    of this specication with gt =q

    j=1 1jHn(tj; ; c)

    2tj.

    Another way of parameterizing the alternative is to assume that the transition variable

    has a xed delay. This assumption results in the following conditional variance model:

    gt = 0d Hn(td; ; c) +q

    j=1

    1j Hn(td; ; c)2tj : (4.4)

  • 8/3/2019 Evaluating GARCH Models

    8/19

    424 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    This specication resembles the one in Terasvirta (1994) for the STAR-type conditional

    mean. Probably the most common case in practice, however, is d = q = p = 1 so that

    (4.3) and (4.4) coincide. Note that we do not impose any nonlinear structure on the

    htj ; j = 1; : : : ; p, as the alternative model is already very exible without such anextension.

    These smooth transition alternatives pose an identication problem. The null hypo-

    thesis can be expressed as H0: = 0 in Hn. It is seen that the remaining parameters

    in (4.2) or (4.3) are only identied under the alternative. In other words, the ele-

    ments of ?2 in gt are nuisance parameters that do not appear in the null model. Thus,

    the classical test statistics are not available as their asymptotic null distributions are

    unknown; see, for example, Hansen (1996) for discussion. We circumvent the identi-

    cation problem by following Luukkonen et al. (1988a). This is done by expanding

    the transition function Hn

    into a rst-order Taylor series around = 0, replacing the

    transition function with this Taylor approximation in (4.3) and rearranging terms. As

    a result,

    ht = (W)st and gt = R

    1v1t +

    n+2i=3

    Ri vit + R1 (4.5)

    in (4.1) and, furthermore, Ri = (i1; : : : ; iq) = Ri; Ri = 0; vit = (it1; : : : ; itq); i =

    1; 3; : : : ; n + 2, and R1 is the remainder. The new null hypothesis, implied by (4.5),

    is H0: R1 = R3 = = Rn+2 = 0. Note that under H0:R1 = 0, so that the stochasticremainder does not aect the distributional properties of the test statistic when thenull hypothesis holds. With ht and gt dened as in (4.5), Eq. (4.1) may be called an

    auxiliary GARCH equation. We make the following additional assumption:

    (A.1) Under H0; E2(n+2)t .

    Remark. The restrictions Assumption (A.1) implies on the parameters in the standard

    GARCH(1; 1) model appeared in Nelson (1990); see also He and Terasvirta (1999).

    Recently; Carrasco and Chen (2002) worked out conditions for the existence of mo-

    ments of higher order than four for the GARCH(p; q) model.

    We dene the conditional quasilog-likelihood for observation t as follows:

    t = (1=2)ln(ht + gt) 2t

    2(ht + gt): (4.6)

    An LM test for the null hypothesis may now be derived starting from (4.6) along the

    same lines as in Appendix A. Note that gt is not a function of W and that under

    H0; gt 0. Here, xt = h1t 9ht=9W (9ht=9W denotes 9(ht + gt)=9W evaluated underH0). In fact, xt has exactly the same form as in Theorem 3.1 because the null model

    is the GARCH(p; q) model in both cases. On the other hand, vt = (v

    1t; v

    3t; : : : ; v

    n+2; t).

    Finally, Assumption (A.1) is needed for the existence of Evn+2; tv

    n+2; t.

    Theorem 4.1. Consider the auxiliary GARCH equation (4.1) where {zt} is a sequenceof independent identically distributed random variables with zero mean; unit variance

  • 8/3/2019 Evaluating GARCH Models

    9/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 425

    and Ez3t = 0; with ht and gt are dened as in (4.5). Under the assumption (A.1) and

    H0: R1 = R3 = = Rn+2 = 0 in (4.5); the statistic (3.3) with xt as in Theorem 3.1 andvt = (v

    1t; v

    3t; : : : ; v

    n+2; t) with vit = (

    it1; : : : ;

    itq)

    ; i = 1; 3; : : : ; n + 2; has an asymptotic

    2 distribution with m = (n + 1)q degrees of freedom.

    A similar result may be derived for the xed delay case. It is also possible to test

    subhypotheses. For example, if the alternative to GARCH is the modied quadratic

    GARCH model in which only the intercept displays a smooth transition, the null hy-

    pothesis equals R1 = 0, whereas R3 = = Rn+2 = 0 even in the maintained model.This test is a linearity test whose alternative generally implies asymmetry. Note,

    however, that if n = 2 and c1 = c2 in (4.2), the smooth transition GARCH modelis still symmetric. Such a model may be useful, for example, in situations where the

    large shocks die out so much more quickly than small shocks that the response cannot be adequately characterized by linear GARCH models.

    The test may be carried out in the TR2 form via an auxiliary regression exactly

    as in the previous section. The same is true for the robust version of the linearity

    test.

    4.3. Testing parameter constancy

    Testing parameter constancy is important in its own right but also because noncon-

    stancy may manifest itself as an apparent lack of covariance stationarity (IGARCH);

    see, for example, Lamoureux and Lastrapes (1990). Here, we assume that the alter-native to constant parameters in the conditional variance is that the parameters, or a

    subset of them, change smoothly over time. Lin and Terasvirta (1994) applied this idea

    to testing parameter constancy in the conditional mean. Let the time-varying parameter

    W(t) =W+[ Hn(t; ; c). If the null hypothesis only concerns a subset of parameters then

    only the corresponding elements in [ are assumed to be nonzero a priori. Again, we

    dene Hn =Hn 12 where the transition function Hn(t; ; c) is assumed to be a logisticfunction of order n dened in (4.2) with xt t. When ; H1(t; ; c) becomes astep-function and characterizes a single structural break in the model. Chu (1995) and

    Lin and Yang (1999) discussed testing parameter constancy against this alternative.

    The null hypothesis of parameter constancy becomes H0: = 0 under which W(t) W.By dening ?1 = and ?2 = ([

    ; c) we can consider this as a special case of (4.1)

    with gt = ([st) Hn(t; ; c). We can again circumvent the lack of identication under

    the null hypothesis by a Taylor approximation of the transition function. A rst-order

    Taylor-expansion of Hn(t; ; c) around = 0 yields, after a reparameterization, model

    (4.1) with

    ht = (W)st and gt =

    ni=1

    Ri vit + R2: (4.7)

    In (4.7), Ri = Ri; Ri = 0; i = 1; : : : ; n. The null hypothesis based on (4.7) equals R1 = = Rn = 0, and the remainder R2 = 0 when the null hypothesis is valid. Furthermore,vit = t

    ist; i = 1; : : : ; n. Dene vit = ti st; i = 1; : : : ; n, and vt = (v

    1t; : : : ; v

    nt).

  • 8/3/2019 Evaluating GARCH Models

    10/19

    426 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    The conditional quasilog-likelihood has the form (4.6), and we can now derive an

    LM test for parameter constancy using the denitions in (4.7). This yields the following

    result:

    Theorem 4.2. Consider the auxiliary GARCH equation (4.1) with (4.7) where {zt}is a sequence of independent identically distributed random variables with zero mean;

    unit variance andEz3t = 0. Under assumption (A.2) and H0: R1 = = Rn = 0 in (4.7);the statistic (3.3) with xt as in Theorem 3.1 and vt=(v1t; : : : ; vnt)

    ; vit=ti st; i =1; : : : ; n ;

    has an asymptotic 2 distribution with m = n(p + q + 1) degrees-of-freedom.

    Remark 1. The proof requires the assumption Ests

    t ; which is equivalent to assum-ing E4t . Since the regularity conditions assumed throughout require even higherorder moments; we do not explicitly mention this condition in the theorem.

    Remark 2. A number of terms in the auxiliary GARCH equation now contain trending

    variables. Nevertheless; applying the results of Lai and Wei (1982) as when proving a

    corresponding result for linear conditional mean models; see Lin and Terasvirta (1994);

    it can be shown that the asymptotic null distribution even in this case is a chi-squared

    one. This makes the test easy to apply.

    Even this test may be carried out using an auxiliary regression as in Section 3.1.

    The procedure for carrying out the robust test, described in Section 3.1, is also valid

    for the parameter constancy test.An advantage of a parametric alternative to parameter constancy is that if the null

    hypothesis is rejected we can estimate the parameters of the alternative model. This

    helps us nd out where in the sample the parameters under test have been changing and

    how rapid the change appears to be. This is useful information when respecication of

    the model to achieve parameter constancy is attempted.

    5. Simulation experiment

    The above distribution theory is asymptotic, and we have to nd out how our tests behave in nite samples. This is done by simulation. For all simulations we use the

    following data generating process (DGP)

    yt = t;

    where t follows a standard GARCH process (size simulations) or one of the alter-

    natives discussed above (power). The random numbers are generated by the random

    number generator in GAUSS 3.2.31. The rst 200 observations of each generated series

    are always discarded to avoid initialization eects. All experiments are performed with

    series of 1000 observations. For each design, a total of 1000 replications have been

    carried out. The distribution for the error process {zt} is either standard normal or astandardized (unit variance) t(d) distribution, where d = 3; 4; 5. We report results from

    experiments with normal and t(3)-errors. The ones with the other two t-distributions

  • 8/3/2019 Evaluating GARCH Models

    11/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 427

    lie between these two extremes. Note that for t(3) and t(4)-errors, the fourth moment

    of t does not exist, which contradicts our assumptions. The idea is to investigate the

    eect of this violation of the conditions on the results.

    Both the normality-based LM and the robust version of each test are considered.This is also true for the tests to which we compare our tests. A general result valid for

    practically all simulation experiments is that the nonrobust test is undersized when the

    error distribution is any one of the three t-distributions. Note that the results appearing

    in the gures are based on size-adjusted tests, so that this nding is not visible there.

    The robust tests are more powerful than the original LM or LM type tests. A general

    conclusion from our experiments is that in applied work one should apply the robust

    versions of our tests.

    5.1. No remaining ARCH

    First, we consider the test of no ARCH in the standardized errors. We dene a DGP

    such that the conditional variance follows a GARCH(1; 2) process:

    ht = 0:5 + 0:052t1 +

    2t2 + 0:9ht1: (5.1)

    In the Monte Carlo experiment the values of vary within limits such that the condi-

    tional variance of the process remains positive with probability 1, see Nelson and Cao

    (1992), and covariance stationarity holds. This is the case when 0:045 0:05.For = 0 the DGP reduces to a standard GARCH(1; 1) model. The results for the test

    of no ARCH in the standardized errors are reported in Fig. 1. The nominal signicance

    level equals 0.1 (the results for levels 0.05 and 0.01 are not reported), and we use

    the test of Bollerslev (1986) against the GARCH(1; 2) model as a benchmark. In the

    simulations, it was always assumed that the additional ARCH component in our test

    was ARCH(1), so that vt= 2t1=ht1. Robust and nonrobust versions of both tests were

    simulated.

    Our results indicate that the robust tests are well-sized, whereas the nonrobust ones

    are undersized when the error distribution is a t-distribution. The size-adjusted power

    results appearing in Fig. 1 are somewhat ambiguous in the normal case. The nonrobust

    tests are more powerful than the robust ones for negative values of but less powerfulfor positive values of this parameter. This is true for both Bollerslevs tests and ours.

    When the errors are t-distributed, a clear dierence in power in favour of the robust

    tests emerges. Note, however, that in that case both tests appear somewhat biased. The

    two tests in our comparison seem to have very similar size-adjusted power, although

    our test has somewhat greater power when the errors follow a t(3)-distribution. As

    Bollerslevs test is an LM test against the data-generating alternative, we may conclude

    that our test works very well in this experiment.

    5.2. Testing linearity

    Our linearity test can be expected to be powerful against smooth transition alterna-

    tives for which it is designed. Assessing its performance against, say, the GJRGARCH

  • 8/3/2019 Evaluating GARCH Models

    12/19

    428 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    Fig. 1. Results from size and power simulations of the LM test of no remaining ARCH (circle) and the

    LM test of GARCH(1,1) against GARCH(1,2) (cross). DGP: GARCH(1,2), Eq. (5.1). x-axis: , y-axis:

    size-adjusted power. Upper panel, left: normal errors, nonrobust test, upper panel, right: normal errors,

    robust test, lower panel, left: t(3)-errors, nonrobust test, lower panel, right: t(3)-errors, robust test. Number

    of observations=1000, number of replications=1000.

    model would constitute a tougher trial for the test. This would also make it possible

    to compare the performance of the tests with the sign-bias and negative size-bias tests

    of Engle and Ng (1993). The DGP is, accordingly, (2.2) withht = 0:005 + 0:23[|t1| + !t1]2 + 0:7ht1; (5.2)

    where t is assumed conditionally normal. In Eq. (5.2), |!| 0:551 is required forcovariance stationarity. This is the model Engle and Ng (1993) used in their simulation

    except that the coecient of [|t1| + !t1]2 equals 0.23 against 0.28 in Engle andNg (1993). This change guarantees the existence of the unconditional fourth moment

    under the null hypothesis ! = 0; the relevant moment condition appeared in He and

    Terasvirta (1999). (The sign-bias and size-bias tests require a nite fourth moment.)

    The sign-bias and size-bias tests are designed for detecting asymmetry in the conditional

    variance. We compute the values of these test statistics using the quadratic form. Thedierence in results between this test and the TR2 version, suggested by Engle and Ng

    (1993), is negligible at our sample size. Engle and Ng (1993) used (2.2) and (5.2)

  • 8/3/2019 Evaluating GARCH Models

    13/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 429

    Fig. 2. Results from size and power simulations of the LM-type parameter constancy test (circle) and the

    robust version of the parameter constancy test of Chu (1995) (cross). DGP: GARCH(1,1) with a single

    structural break, equation (5.3). x-axis: , y-axis: size-adjusted power. Upper panel, left: normal errors,

    nonrobust test, upper panel, right: normal errors, robust test, lower panel, left: t(3)-errors, nonrobust test,

    lower panel, right: t(3)-errors, robust test. Number of observations=1000, number of replications=1000.

    with !=

    0:23 as the DGP in their evaluation of the sign-bias test. We also considered

    robust versions of the two tests.The results for the sign-bias and negative size-bias tests are rather similar. Thus, the

    latter test is excluded from Fig. 2 where our test is compared with the sign-bias test.

    The nonrobust sign-bias and our linearity test are undersized for the t(3)-errors. The

    robust versions of the tests have roughly the correct size.

    As for size-adjusted power, results for ! 0 can be found in Fig. 2. For positive

    values of ! the results are similar and therefore omitted. Our smooth transition GARCH

    test was computed by assuming n = 1 in (4.2). When the errors are normal, 1000

    observations are enough for the robust tests to perform as well as the nonrobust ones.

    In the case of t-distributed errors, the robust tests are always more powerful than

    the nonrobust ones. Our linearity test is clearly more powerful than the sign-bias testeven if the alternative is a GJRGARCH and not a genuine smooth transition GARCH

    model.

  • 8/3/2019 Evaluating GARCH Models

    14/19

    430 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    Fig. 3. Results from power simulations of the LM-type parameter constancy test (circle) and the parameter

    constancy test of Chu (1995) (cross). DGP: GARCH(1,1) with a double structural break, Eq. (5.3). x-axis:

    , y-axis: size-adjusted power. Upper panel, left: normal errors, nonrobust test, upper panel, right: normal

    errors, robust test, lower panel, left: t(3)-errors, nonrobust test, lower panel, right: t(3)-errors, robust test.

    Number of observations=1000, number of replications=1000.

    5.3. Testing parameter constancy

    We consider two cases of parameter nonconstancy: the DGP is a GARCH(1; 1)

    model with either (a) a single or (b) a double structural break. We did not simulate

    smooth parameter change because our test can be expected to perform well against

    such an alternative. Our choice of alternative gives us an opportunity to compare our

    test with tests against a single structural break, and we choose the test of Chu (1995)

    for the purpose. If T is the total number of observations, the single structural break

    parameterization (a) is assumed to have a break at time T where lies between 0

    to 1. The double structural break parameterization (b) postulates a change at time 1T

    and a return to the original parameters at 2T; 06 1 26 1. Our test is computedwith n = 1, case (a), and n = 2, case (b), where n is the order of the logistic function

    in (4.2).

  • 8/3/2019 Evaluating GARCH Models

    15/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 431

    Fig. 4. Results from size and power simulations of the LM-type linearity test (circle) and the sign-bias test

    of Engle and Ng (1993) (cross). DGP: GJRGARCH(1,1), Eq. (5.2). x-axis: !, y-axis: size-adjusted power.

    Upper panel, left: normal errors, nonrobust test, upper panel, right: normal errors, robust test, lower panel,

    left: t(3)-errors, nonrobust test, lower panel, right: t(3)-errors, robust test. Number of observations=1000,

    number of replications=1000.

    We consider the following model for a change in the constant term:

    ht = 0:5 + 0:12t1 + 0:8ht1; (a) t T; (b) t 1T; t 2T;

    ht = 0:5(1 + ) + 0:12t1 + 0:8ht1; (a) t T; (b) 1T6 t6 2T; (5.3)

    where = 0:4; 0:8. Chu (1995) used (a) in (5.3) as the DGP in his own simulation

    experiments. The power simulations for the DGPs in (5.3) with a single structural

    break at for = 0:8 appear in Fig. 3. The values = 0; 1 correspond to the null

    hypothesis. Even if only the intercept changes in (5.3), in simulating our LM-type

    test we assume that under the alternative, the break aects all three parameters. The

    asymptotic null distribution of our LM-type statistic is thus the 2(3)-distribution.

    We report results for = 0:8. The pattern of the results for = 0:4 is similar, only

    the power is lower. In another experiment, we allowed the coecient of 2t1 to changeonce within the sample period. The behaviour of the tests was similar to the previous

    case, and no details are given here. In case (a) under normality, our LM-type test

  • 8/3/2019 Evaluating GARCH Models

    16/19

    432 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    has the same as or higher power than the test of Chu (1995) for 0:36 6 0:7, but

    otherwise the relationship is the opposite. When the errors follow a t(3)-distribution,

    the robust version of our test is more powerful than the standard version. It is also again

    more powerful than Chus test for 0:36 6 0:7. It seems, however, that Chus robusttest is oversized when the errors are t(3)-distributed whereas our robust one is not.

    On the other hand, in that situation the nonrobust LM-type test is clearly undersized

    (these two facts cannot be seen from Fig. 3).

    We turn to case (b), the double structural break. The DGP in that design is such

    that 2 = 1 + 0:3 where 1 varies from 0 to 0:7. Thus, for 1 =0 and 1 = 0:7, the DGP

    only has a single structural break. Results of power simulations of this design with

    = 0:8 can be found in Fig. 4. In the case of a double break the test of Chu (1995)

    cannot be expected to be very powerful because the test is constructed for detecting

    just a single structural break. For this reason, its power is higher for 1 close to zero

    and 0:7 than elsewhere. With few exceptions (nonrobust test, t(3)-distributed errors)

    our test with n =2 does have size-adjusted power superior to that of Chu for all double

    break points considered. For t-distributed errors, the robust version of the test is clearly

    more powerful than the nonrobust one.

    6. Conclusions

    In this paper, we derive a number of misspecication tests for a standard GARCH

    (p; q) model. As all our test statistics are asymptotically 2-distributed under the null

    hypothesis, possible misspecication of a GARCH model can be detected at low com- putational cost. Because the tests of linearity and parameter constancy are parametric,

    the alternative may be estimated if the null hypothesis is rejected. This is useful for

    a model builder who wants to nd out the possible weaknesses of the estimated spec-

    ication. It may also give him=her useful ideas of how the model could be further

    improved.

    We also highlight the fact that our test of no ARCH in the standardized error process

    is asymptotically equivalent to a portmanteau test of Li and Mak (1994). This links

    the work of these authors to our approach. As already mentioned, the advantage of our

    derivation of the no ARCH test statistic is that robustifying the test against nonnormal

    errors is straightforward.

    Finally, the simulation results indicate that in practice, the robust versions of the tests

    should be preferred to nonrobust ones. At relevant sample sizes when the errors are

    normal, they are roughly as powerful as normality-based LM or LM type tests. When

    the errors are nonnormal, the robust tests are superior in terms of power to nonrobust

    ones. They can therefore be recommended as standard procedures when it comes to

    testing the adequacy of an estimated standard GARCH(p; q) model.

    Acknowledgements

    The rst author acknowledges nancial support from the Tore Browaldhs Foundation

    and Skandia Asset Management. The research of the second author has been supported

  • 8/3/2019 Evaluating GARCH Models

    17/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 433

    by the Swedish Council for Research in Humanities and Social Sciences. We thank

    James Davidson and an anonymous referee for helpful suggestions and remarks but

    retain the responsibility for any errors and shortcomings in the paper.

    Appendix A. Proof of Theorem 3.1

    The conditional quasi log likelihood for observation t equals

    t = 12

    ln ht 12

    ln gt 2t

    2htgt:

    The partial derivative with respect to W is

    9t

    9W= 1

    2

    1

    ht

    9ht

    9W+

    1

    gt

    9gt

    9W

    2t

    h2tgt

    1

    ht

    9ht

    9W

    2t

    htg2t

    1

    gt

    9gt

    9W

    =1

    2(2t =htgt 1)xt +

    1

    2(2t=htgt 1)

    1

    gt

    9gt

    9W;

    where xt = h1t 9ht=9W with 9ht=9W = st + (9st=9Wt)

    W, and 9gt=9W = (9vt=9Wt)?. Thus,

    9gt=9W|H0 = 0 so that 9t=9W|H0 = ( 12 )(2t=ht 1)xt. Likewise,

    9t9?

    = 12

    (2t=htgt 1) 1gt 9gt

    9?

    so that 9t=9?|H0 = ( 12 )(2t=ht 1)vt, where vt = (2t1=ht1; : : : ; 2tm=htm). Let G =(D; W). The relevant component of the average pseudoscore evaluated at the true

    (under the null hypothesis) pair of parameter values (G0; 0) has the form

    qT(G0; 0) =1

    2T

    Tt=1

    (2t=ht 1)

    x0t

    vt

    ;

    where x0t equals xt evaluated at (G0; 0). (We need not consider (1=T)

    Tt=1 9t=9D|H0 ,

    because assuming Ez3t = 0 makes the covariance matrix of the full pseudoscore vector

    block diagonal.) Let t = yt f(wt; D) where D is the quasi maximum likelihoodestimator ofD under H0. Furthermore, let ht; xt and vt equal ht; xt and vt evaluated at

    G = G. Under regularity conditions,

    TqT(G; 0) =

    1

    2

    T

    Tt=1

    (2t=ht 1)

    xtvt

    =

    0

    (1=2

    T)

    Tt=1

    (2t=ht 1)vt N(0; J0);

  • 8/3/2019 Evaluating GARCH Models

    18/19

    434 S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435

    in distribution as T , where

    J0 = J011 J012J021 J022

    = ET1qT(G0; 0)qT(G0; 0)

    =k

    4E

    x0t x

    0

    t x0t v

    t

    x0tv

    t vtv

    t

    :

    with k = E(2t=ht 1)2. Thus,

    LM0 =

    T

    t=1

    (2t=ht 1)vt)J1022

    T

    t=1

    (2t=ht 1

    vt

    (A.1)

    has an asymptotic 2

    distribution with m degrees of freedom. Replacing J1

    022 in (A.1) by a consistent estimator

    V(W)1 = (4T=k)

    Tt=1

    vtv

    t T

    t=1

    vtx

    t

    T

    t=1

    xtx

    t

    1 Tt=1

    xtv

    t

    1

    with k= (1=T)T

    t=1(2t=ht1)2, yields the result; see Davidson (2000, Section 12.3.3).

    References

    Anderson, H.M., Nam, K., Vahid, F., 1999. Asymmetric nonlinear smooth transition GARCH models. In:

    Rothman, P. (Ed.), Nonlinear time series analysis of economic and nancial data. Kluwer, Boston.

    Bollerslev, T., 1986. Generalized autoregressive conditional heteroskedasticity. Journal of Econometrics 31,

    307327.

    Carrasco, M., Chen, X., 2002. Mixing and moment properties of various GARCH and stochastic volatility

    models. Econometric Theory 18, 1739.

    Chu, C.-S.J., 1995. Detecting parameter shift in GARCH models. Econometric Reviews 14, 241266.

    Comte, F., Lieberman, O., 2000. Asymptotic theory for multivariate GARCH processes. Journal of

    Multivariate Analysis, forthcoming.

    Davidson, J., 2000. Economic Theory. Blackwell, Oxford.

    Demos, A., Sentana, E., 1998. Testing for GARCH eects: a one-sided approach Journal of Econometrics

    86, 97127.

    Eitrheim, I., Terasvirta, T., 1996. Testing the adequacy of smooth transition autoregressive models. Journal

    of Econometrics 74, 5975.

    Engle, R.F., 1982. Autoregressive conditional heteroscedasticity with estimates of the variance of United

    Kingdom ination. Econometrica 50, 9871007.

    Engle, R.F., Ng, V.K., 1993. Measuring and testing the impact of news on volatility. Journal of Finance 48,

    17491777.

    Glosten, L., Jagannathan, R., Runkle, D., 1993. On the relation between expected value and the volatility of

    the nominal excess return on stocks. Journal of Finance 48, 17791801.

    Gonzalez-Rivera, G., 1998. Smooth transition GARCH models. Studies in Nonlinear Dynamics and

    Econometrics 3, 6178.Hagerud, G.E., 1997. A New Non-Linear GARCH Model. EFI, Stockholm School of Economics.

    Hansen, B.E., 1996. Inference when a nuisance parameter is not identied under the null hypothesis.

    Econometrica 64, 413430.

  • 8/3/2019 Evaluating GARCH Models

    19/19

    S. Lundbergh, T. Terasvirta / Journal of Econometrics 110 (2002) 417 435 435

    He, C., Terasvirta, T., 1999. Properties of moments of a family of GARCH processes. Journal of

    Econometrics 92, 173192.

    Lai, T.Z., Wei, C.Z., 1982. Least squares estimates in stochastic regression models with applications in

    identication and control of dynamic systems. Annals of Statistics 10, 154166.Lamoureux, C.G., Lastrapes, W.G., 1990. Persistence in variance, structural change and the GARCH model.

    Journal of Business and Economic Statistics 8, 225234.

    Lee, J., King, M., 1993. A locally most mean powerful based score test for ARCH and GARCH regression

    disturbances. Journal of Business and Economic Statistics 11, 1727.

    Li, W.K., Mak, T.K., 1994. On the squared residual autocorrelations in non-linear time series with conditional

    heteroskedasticity. Journal of Time Series Analysis 15, 627636.

    Lin, C.-F., Terasvirta, T., 1994. Testing the constancy of regression parameters against continuous structural

    change. Journal of Econometrics 62, 211228.

    Lin, S.-J., Yang, J., 1999. Testing shift in nancial models with conditional heteroskedasticity: an empirical

    distribution function approach. Research Paper 30, University of Technology, Quantitative Finance

    Research Group, Sydney.

    Luukkonen, R., Saikkonen, P., Terasvirta, T., 1988a. Testing linearity against smooth transition autoregressivemodels. Biometrika 75, 491499.

    Luukkonen, R., Saikkonen, P., Terasvirta, T., 1988b. Testing linearity in univariate time series models.

    Scandinavian Journal of Statistics 15, 161175.

    McLeod, A.I., Li, W.K., 1983. Diagnostic checking ARMA time series models using squared residual

    autocorrelations. Journal of Time Series Analysis 4, 269273.

    Nelson, D.B., 1990. Stationarity and persistence in the GARCH(1,1) model. Econometric Theory 6, 318334.

    Nelson, D.B., Cao, C.Q., 1992. Inequality constraints in the univariate GARCH model. Journal of Business

    and Economic Statistics 10, 229235.

    Sentana, E., 1995. Quadratic ARCH models. Review of Economic Studies 62, 639661.

    Terasvirta, T., 1994. Specication, estimation, and evaluation of smooth transition autoregressive models.

    Journal of the American Statistical Association 89, 208218.

    Wooldridge, J., 1991. On the application of robust, regression-based diagnostics to models of conditional

    means and conditional variances. Journal of Econometrics 47, 546.